Phosphaspiropentene as a Transient Intermediate - Organometallics

J. Chris Slootweg, Willem-Jan van Zeist, Frans J. J. de Kanter, Marius Schakel, Andreas W. Ehlers, Martin Lutz, Anthony L. Spek, and Koop Lammertsma*...
0 downloads 0 Views 123KB Size
5172

Organometallics 2005, 24, 5172-5175

Notes Phosphaspiropentene as a Transient Intermediate J. Chris Slootweg,† Willem-Jan van Zeist,† Frans J. J. de Kanter,† Marius Schakel,† Andreas W. Ehlers,† Martin Lutz,‡ Anthony L. Spek,‡ and Koop Lammertsma*,† Department of Chemistry, Faculty of Sciences, Vrije Universiteit, De Boelelaan 1083, 1081 HV Amsterdam, The Netherlands, and the Bijvoet Center for Biomolecular Research, Crystal and Structural Chemistry, Utrecht University, Padualaan 8, 3584 CH Utrecht, The Netherlands Received June 27, 2005 Summary: A phosphaspiropentene is the plausible kinetic product from the addition of dichlorocarbene to a phosphatriafulvene, which rearranges to a novel Psubstituted triafulvene. The calculated barrier of 18.6 kcal mol-1 for this process is consistent with the temperature of -40 °C at which this reaction proceeds. Introduction Spiro-connected cycloalkenes exhibit intriguing bonding and electronic properties due to spiroconjugation.1 Although the smallest of these highly strained hydrocarbons, spiropentene (1)2 and spiropentadiene (2),3

have been reported, spirenes with heteroatoms are extremely rare. Only recently, the thermally stable spiropentasiladiene 3 was reported.4 Examples with other heteroatoms are limited to 1-oxaspiropent-4-enes, including the parent 4,5 and 1-azaspiropent-1-ene 5,6 but none with a phosphorus atom are known other than * To whom correspondence should be addressed. Tel: +31-205987474. Fax: +31-20-5987488. E-mail: [email protected]. † Vrije Universiteit. ‡ Utrecht University. (1) Liebman, J. F.; Greenberg, A. Chem. Rev. 1976, 76, 311-365. (2) (a) Bloch, R.; Denis, J. M. Angew. Chem. 1980, 92, 969-970; Angew. Chem., Int. Ed. Engl. 1980, 19, 928-929. (b) Irngartinger, H.; Gries, S. Angew. Chem. 1991, 103, 595-596; Angew. Chem., Int. Ed. Engl. 1991, 30, 565-566. (c) Wade, P. A.; Kondracki, P. A. J. Chem. Soc., Chem. Commun. 1994, 1263-1264. (3) (a) Billups, W. E.; Haley, M. H. J. Am. Chem. Soc. 1991, 113, 5084-5085. (b) Saini, R. K.; Litosh, V. A.; Daniels, A. D.; Billups, W. E. Tetrahedron Lett. 1999, 40, 6157-6158. (4) Iwamoto, T.; Tamura, M.; Kabuto, C.; Kira, M. Science 2000, 290, 504-506.

the 1-aza-2-phosphaspiro[2.2]pentene 6, which may be a transient in the formation of 1H-2-iminophosphetes.7 In contrast, P-containing spiranes carrying a transitionmetal complex, e.g. the phosphaspiropentane 7,8 are stable, which is highlighted for the extended arrays9 by a phospha[7]triangulane that consists of seven spiroconnected three-membered rings and has a melting point above 150 °C.10 Results and Discussion Can a transition-metal complex likewise stabilize a P-containing spirene? To explore this, we examine the chemistry of 1-phosphaspiropent-4-ene. Access to this compound can be envisioned by addition of dichlorocarbene, generated in situ from t-BuOK/CHCl3,11 to the exocyclic PdC bond12 of the phosphatriafulvene complex 813 (0 °C, pentane). This reaction yielded, instead, the novel P-substituted triafulvene 10 (95%, colorless crystals) and showed no trace of phosphaspiropentene 9, not even by 31P NMR monitoring of the reaction at -40 °C (Scheme 1). Triafulvene 10 has distinctive resonances at δ(31P) 110.0 (1J(P,W) ) 282.7 Hz) and δ(13C) 86.6 (1J(C,P) ) 52.5 Hz), 135.2 (d, 2J(C,P) ) 9 Hz), 146.0 (s), and 147.4 (3J(C,P) ) 6.4 Hz).14 P-C bond rotation of the mesityl group is hindered, causing a broad signal (5) (a) Billups, W. E.; Litosh, V. A.; Saini, R. K.; Daniels, A. D. Org. Lett. 1999, 1, 115-116. (b) Bickers, P. T.; Halton, B.; Kay, A. J.; Northcote, P. T. Aust. J. Chem. 1999, 52, 647-652. (c) Halton, B.; Cooney, M. J.; Wong, H. J. Am. Chem. Soc. 1994, 116, 11574-11575. (6) (a) Irngartinger, H.; Gries, S. Chem. Ber. 1992, 125, 2513-2515. (b) A 1,2-diazaspiro[2.2]pent-1-ene was also reported: Patent, Bergwerksverband G.m.b.H., GB 1097238, 1965. (7) Eisfeld, W.; Slany, M.; Bergstra¨sser, U.; Regitz, M. Tetrahedron Lett. 1994, 35, 1527-1530. (8) Hung, J.-T.; Yang, S.-W.; Gray, G. M.; Lammertsma, K. J. Org. Chem. 1993, 58, 6786-6790. (9) Slootweg, J. C.; de Kanter, F. J. J.; Schakel, M.; Lutz, M.; Spek, A. L.; Kozhushkov, S. I.; de Meijere, A.; Lammertsma, K. Chem., Eur. J., in press. (10) Slootweg, J. C.; Schakel, M.; de Kanter, F. J. J.; Ehlers, A. W.; Kozhushkov, S. I.; de Meijere, A.; Lutz, M.; Spek, A. L.; Lammertsma, K. J. Am. Chem. Soc. 2004, 126, 3050-3051. (11) Doering, W. v. E.; Hoffmann, A. K. J. Am. Chem. Soc. 1954, 76, 6162-6165. (12) Yoshifuji, M.; Toyota, K.; Yoshimura, H.; Hirotsu, K.; Okamoto, A. J. Chem. Soc., Chem. Commun. 1991, 124-125 and references therein. (13) Fuchs, E.; Breit, B.; Heydt, H.; Schoeller, W.; Busch, T.; Kru¨ger, C.; Betz, P.; Regitz, M. Chem. Ber. 1991, 124, 2843-2855.

10.1021/om050530+ CCC: $30.25 © 2005 American Chemical Society Publication on Web 09/15/2005

Notes

Organometallics, Vol. 24, No. 21, 2005 5173

Figure 1. Displacement ellipsoid plot of 10 with ellipsoids drawn at the 50% probability level. Hydrogen atoms and the pentane solvent molecule are omitted for clarity. Selected bond lengths (Å), bond angles (deg), and torsion angles (deg): W1-P1 ) 2.4987(7), Cl1-P1 ) 2.0920(10), Cl2-C1 ) 1.774(3), P1-C1 ) 1.782(3), P1-C5 ) 1.837(3), C1-C2 ) 1.331(4), C2-C3 ) 1.433(4), C2-C4 ) 1.425(4), C3-C4 ) 1.331(4); Cl2-C1-P1 ) 115.04(15), C2-C3-C4 ) 62.0(2), C3-C2-C4 ) 55.49(19), C2-C4-C3 ) 62.5(2); C3-C2-C1-P1 ) -1.4(8). Scheme 1. Synthesis of Triafulvene 10

Scheme 2. Rearrangements of Phosphirane 11

at δ(1H, 293 K) 2.54 for the o-methyl groups that narrows at higher temperatures. The single-crystal X-ray structure (Figure 1) shows a shortened P-C1 bond (1.782(3) Å) and the cyclopropene ring being nearly coplanar with the Cl2-C1-P1 plane with a C3-C2C1-P1 torsion angle of -1.4(8)°. Normal CdC bonds (C1-C2 ) 1.331(4) Å, C3-C4 ) 1.331(4) Å) indicate a diminished π delocalization, which is confirmed by the calculated NICS value of only -21.4, which is considerably less negative than that of cyclopropene (-28.4). Is phosphaspiropentene 9 an intermediate in the formation of 10? Such a rearrangement does convert the phosphirane complex 11 into 12 (Scheme 2), but at the much higher temperature of 110 °C.15 We examine both processes using theoretical methods. A large barrier of 71.3 kcal mol-1 (B3LYP/6-31G**) has been reported for converting the parent 11′ into 12′ (H for Me, Ph; no W(CO)5) by P-C bond cleavage with a concurrent Cl shift from C to P,16 which, of course, does not comply with the experimental observations,15 nor does the calculated favored H shift that would give 13. Clearly, the transition-metal group has a major influence, which we substantiate using BP86/6-31G**(LANL2DZ) calculations.17 (14) Schubert, H. H.; Stang, P. J. J. Org. Chem. 1984, 49, 50875090. (15) Tran Huy, N. H.; Mathey, F. J. Org. Chem. 2000, 65, 652654. (16) Ma´trai, J.; Dransfield, A.; Veszpre´mi, T.; Nguyen, M. T. J. Org. Chem. 2001, 66, 5671-5678.

Figure 2. Relative BP86/6-31G** (LANL2DZ for W) energies (ZPE corrected, in kcal mol-1) for the conversion of 11 into 12. Selected bond lengths (Å) of 11: W1-P1 ) 2.539, P1-C1 ) 1.877, P1-C2 ) 1.909, P1-C3 ) 1.834, C1-Cl1 ) 1.785, C1-C2 ) 1.539. Selected bond lengths (Å) of TSantiCl-shift (TS11 f 12): W1-P1 ) 2.475, P1Cl1 ) 3.508, P1-C1 ) 1.782, P1-C2 ) 2.751, P1-C3 ) 1.806, C1-Cl1 ) 2.313, C1-C2 ) 1.418. Selected bond lengths (Å) of 12: W1-P1 ) 2.527, P1-Cl1 ) 2.133, P1C1 ) 1.832, P1-C3 ) 1.846, C1-C2 ) 1.356.

Adding the singlet phosphinidene PhPdW(CO)5 to 1-chloro-2-methylpropene gives 11, likely by a barrierfree non-least-motion trajectory,18 with less exothermicity (23.0 kcal mol-1) than for the parent ethylene (35.9 kcal mol-1),19 due to the reduced nucleophilicity of 1-chloro-2-methylpropene. Converting structure 11 in a single step to the 20.9 kcal mol-1 more stable vinylchlorophosphine 12, in which the chloride anti to the P-W(CO)5 group migrates to phosphorus, requires a 24.0 kcal mol-1 barrier to be overcome (Figure 2). Converting phosphirane 11 into vinylphosphine 13, by migrating H instead of Cl, has only a marginal exothermicity (3.2 kcal mol-1) and a higher barrier (∆Eq ) 39.1 kcal mol-1), suggesting this to be a less likely process, which concurs with the experimental observations. The barrier for rearranging 11 into 12 is comparable to the dissociation energy to regenerate PhPdW(CO)5 and 1-chloro-2-methylpropene (23.0 kcal mol-1), which clarifies the modest isolated yields for 11 (11%) and 12 (39%).15 In contrast, the analogous phosphirane 14 does not rearrange thermally and a Pd(PPh3)4-catalyzed reaction is needed (85 °C) to afford vinylchlorophosphine 15 (Scheme 3).15 BP86/6-31G**(LANL2DZ) calculations reveal that converting structure 14 to the 18.6 kcal mol-1 more stable vinylchlorophosphine 15, in which the (17) Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Zakrzewski, V. G.; Montgomery, J. A., Jr.; Stratmann, R. E.; Burant, J. C.; Dapprich, S.; Millam, J. M.; Daniels, A. D.; Kudin, K. N.; Strain, M. C.; Farkas, O.; Tomasi, J.; Barone, V.; Cossi, M.; Cammi, R.; Mennucci, B.; Pomelli, C.; Adamo, C.; Clifford, S.; Ochterski, J.; Petersson, G. A.; Ayala, P. Y.; Cui, Q.; Morokuma, K.; Malick, D. K.; Rabuck, A. D.; Raghavachari, K.; Foresman, J. B.; Cioslowski, J.; Ortiz, J. V.; Stefanov, B. B.; Liu, G.; Liashenko, A.; Piskorz, P.; Komaromi, I.; Gomperts, R.; Martin, R. L.; Fox, D. J.; Keith, T.; Al-Laham, M. A.; Peng, C. Y.; Nanayakkara, A.; Gonzalez, C.; Challacombe, M.; Gill, P. M. W.; Johnson, B. G.; Chen, W.; Wong, M. W.; Andres, J. L.; Head-Gordon, M.; Replogle, E. S.; Pople, J. A. Gaussian 98, revision A.11.4; Gaussian, Inc.: Pittsburgh, PA, 1998. See the Supporting Information. (18) (a) Bulo, R. E.; Ehlers, A. W.; de Kanter, F. J. J.; Schakel, M.; Lutz, M.; Spek, A. L.; Lammertsma, K.; Wang, B. Chem. Eur. J. 2004, 10, 2732-2738. (b) Wit, J. B. M. Ph.D. Thesis, Vrije Universiteit, Amsterdam, The Netherlands, 2001, Chapter 7. (19) 66 kcal mol-1 at B3LYP/6-31G*: Keating, A. E.; Merrigan, S. R.; Singleton, D. A.; Houk, K. N.; J. Am. Chem. Soc. 1999, 121, 39333938.

5174

Organometallics, Vol. 24, No. 21, 2005

Notes

Scheme 3. Rearrangement of Phosphirane 14

chloride anti to the P-W(CO)5 group migrates to phosphorus,20 is more demanding than the rearrangement of 11 and requires a 31.5 kcal mol-1 barrier to be overcome. This barrier is much higher than the dissociation energy needed to regenerate Ph-PdW(CO)5 and 1,1-dichloroethene (23.2 kcal mol-1), explaining why this reaction does not proceed thermally. Extending the calculations to the addition of CCl2 to phosphatriafulvene 8′′ (H for tBu, Ph for mesityl; labeled ′′) indicates a large exothermicity (52.3 kcal mol-1) for the formation of phosphaspiropentene 9′′. We note that the addition of PhPdW(CO)5 to 3,3-dichlorotriafulvene, which would also give 9′′, is less exothermic (22.9 kcal mol-1), because metal complexation reduces the reactivity of phosphinidenes.21 Without the W(CO)5 group, 1PPh (A ) adds to 3,3-dichlorotriafulvene with a reaction 1 energy of -71.7 kcal mol-1. The single-step conversion of the W(CO)5-complexed phosphaspiropentene 9′′ into chlorophosphine 10′′ occurs more easily (∆Eq ) 18.6 kcal mol-1, ∆E ) -24.7 kcal mol-1; Figure 3)22 than the analogous 14 f 15 rearrangement (∆Eq ) 31.5 kcal mol-1), and is favored by 4.6 kcal mol-1 over the dissociation into PhPdW(CO)5 and 3,3-dichlorotriafulvene. The higher reactivity of 9′′ compared to that of 11 can be rationalized by the destabilization of the phospirene ring upon spiro fusion. Both the phosphorus23 and the electron-withdrawing chloro substituents activate phosphirene 14 (NICS ) -34.8) compared to the parent cyclopropene24 (NICS ) -42.8); additional spiro fusion in 9′′ further reduces the σ aromaticity (NICS ) -26.9). This effect is also reflected by the elongated distal and proximal P-C bonds9 (P1-C1 ) 1.927 Å, P1-C2 ) 1.866 Å) as compared to the corresponding bond lengths of 14 (P1C1 ) 1.897 Å, P1-C2 ) 1.847 Å) and phosphaspiropentane 78 (experimental: P1-C1 ) 1.855 Å, P1-C2 ) 1.794 Å). Conclusions Phosphaspiropentene 9 is the plausible kinetic product that results from addition of dichlorocarbene to phosphatriafulvene 8. The calculated barrier of 18.6 kcal mol-1 for rearrangement to the more stable P-substituted triafulvene 10 is consistent with the temperature of -40 °C at which the reaction proceeds. Currently, we are testing other carbenes to obtain the, as yet, elusive phosphaspiropentenes. (20) The syn Cl shift is less favorable: ∆Eq ) 41.1 kcal mol-1. (21) Ehlers, A. W.; Baerends, E. J.; Lammertsma, K. J. Am. Chem. Soc. 2002, 124, 2831-2838. (22) There are numerous examples of DFT overestimating activation barriers, e.g.: Goumans, T. P. M.; Ehlers, A. W.; Lammertsma, K.; Wu¨rthwein, E.-U.; Grimme, S. Chem., Eur. J. 2004, 10, 6468-6475. (23) Li, Z.-H.; Moran, D.; Fan, K.-N.; Schleyer, P. v. R. J. Phys. Chem. A 2005, 109, 3711-3716. (24) (a) Dewar, M. J. S. Bull Soc. Chim. Belg. 1979, 88, 957-967. (b) Dewar, M. J. S. J. Am. Chem. Soc. 1984, 106, 669-682. (c) Cremer, D.; Kraka, E. J. Am. Chem. Soc. 1985, 107, 3800-3810. (d) Cremer, D.; Gauss, J. J. Am. Chem. Soc. 1986, 108, 7467-7477.

Figure 3. Relative BP86/6-31G** (LANL2DZ for W) energies (ZPE corrected, in kcal mol-1) for the conversion of 9′′ into 10′′. Selected bond lengths (Å) of 9′′: W1-P1 ) 2.520, P1-C1 ) 1.927, P1-C2 ) 1.866, P1-C5 ) 1.834, C1-Cl1 ) 1.789, C1-Cl2 ) 1.791, C1-C2 ) 1.498, C2C3 ) 1.489, C2-C4 ) 1.491, C3-C4 ) 1.312. Selected bond lengths (Å) of TS9′′-10′′: W1-P1 ) 2.497, P1-Cl1 ) 3.525, P1-C1 ) 1.810, P1-C2 ) 2.536, P1-C5 ) 1.826, C1-Cl1 ) 2.239, C1-Cl2 ) 1.783, C1-C2 ) 1.416, C2-C3 ) 1.426, C2-C4 ) 1.420, C3-C4 ) 1.344. Selected bond lengths (Å) of 10′′: W1-P1 ) 2.517, P1-Cl1 ) 2.141, P1-C1 ) 1.814, P1-C5 ) 1.854, C1-Cl2 ) 1.771, C1-C2 ) 1.350, C2-C3 ) 1.442, C2-C4 ) 1.438, C3-C4 ) 1.335.

Experimental Section Computations. All density functional theory calculations (BP86) were performed with the Gaussian98 suite of programs,17 using the LANL2DZ basis and pseudopotentials for tungsten and the 6-31G** basis for all other atoms. The natures of all transition structures were confirmed with frequency calculations. Intrinsic reaction coordinate (IRC) calculations were performed to ascertain the connection between reactant and product. The nucleus independent chemical shift25 (NICS) values were calculated at the B3LYP/6311G+(2p,d) level, leaving out the substituents. General Considerations. NMR spectra were recorded on Bruker Advance 250 (31P; 85% H3PO4) and MSL 400 instruments (1H, 13C) and referenced internally to residual solvent resonances (1H, δ 7.25 ppm (CDCl3); 13C{1H}, 77.0 ppm (CDCl3)). The IR spectrum was recorded on a Mattson-6030 Galaxy FT-IR spectrophotometer, and the high-resolution mass spectrum (HR-MS) was performed on a Finnigan Mat 900 mass spectrometer operating at an ionization potential of 70 eV. The melting point of 10 was measured on a sample in an unsealed capillary and is uncorrected. Synthesis of 10. CHCl3 (56 µL, 0.7 mmol) was added under nitrogen at 0 °C to a solution of 813 (87 mg, 0.14 mmol) and t-BuOK (79 mg, 0.7 mmol) in dry pentane (3 mL). After additional stirring for 30 min at 0 °C and another 30 min at room temperature, the reaction mixture was filtered and concentrated and 10 could be obtained as colorless crystals in 95% yield (104 mg; )10‚(pentane)) after crystallization at 20 °C. Characterization data for 10: mp 100 °C dec; 31P{1H} NMR (101.3 MHz, CDCl3, 293 K) δ 110.0 (1J(P,W) ) 282.7 Hz); 13 C{1H} NMR (100.6 MHz, CDCl3, 328 K) δ 20.5 (s; p-CH3ArP), 24.5 (d, 3J(C,P) ) 5.4 Hz; o-CH3-ArP), 28.4 (s; C(CH3)3), 29.6 (s; C(CH3)3), 32.3 (s; C(CH3)3), 32.4 (d, 4J(C,P) ) 0.8 Hz; C(CH3)3), 86.6 (d, 1J(C,P) ) 52.5 Hz; dCCl), 131.4 (d, 3J(C,P) ) 7.3 Hz; m-ArP), 135.2 (d, 2J(C,P) ) 9 Hz; o-ArP), 135.2 (d, 2 J(C,P) ) 9 Hz; CdCCl), 135.7 (d, 1J(C,P) ) 26.8 Hz; ipsoArP), 139.9 (d, 4J(C,P) ) 1.5 Hz; p-ArP), 146.0 (s; dCC(CH3)3), 147.4 (d, 3J(C,P) ) 6.4 Hz; dCC(CH3)3), 196.8 (d, 2J(C,P) ) 7.3 Hz, 1J(C,W) ) 127.2 Hz; cis-CO), 200.3 (d, 2J(C,P) ) 31.1 Hz, 1J(C,W) ) 141.6 Hz; trans-CO); 1H NMR (400.1 MHz, CDCl3, 328 K) δ 0.86 (s, 9H; C(CH3)3), 1.40 (s, 9H; C(CH3)3), (25) Schleyer, P. v. R.; Maerker, C.; Dransfeld, A.; Jiao, H.; van Eikema Hommes, N. J. R. J. Am. Chem. Soc. 1996, 118, 6317-6318.

Notes 2.25 (s, 3H; p-CH3-ArP), 2.54 (br s, 6H; o-CH3-ArP), 6.84 (d, 4 J(H,P) ) 4.2 Hz, 2H; m-ArP); IR (KBr) ν 1939 (s/br, COeq), 2074 cm-1 (w, COax); MS (EI, 70 eV) m/z (%) 706 (0.5) [M]+, 671 (0.5) [M - Cl]+, 566 (10) [M - 5CO]+; HR-MS m/z calcd for [M - 5CO]+ 566.0893, found 566.0900. Crystal Structure Determination of Compound 10. Crystal data: C26H29Cl2O5PW‚0.5C5H12, fw 743.28, colorless needle, 0.42 × 0.24 × 0.06 mm3; monoclinic crystal system, space group C2/c (No. 15); cell parameters a ) 19.0406(2) Å, b ) 18.9032(2) Å, c ) 19.5695(2) Å, β ) 117.0955(5)°, V ) 6270.57(11) Å3. Z ) 8, F ) 1.575 g/cm3. A total of 71 511 reflections were measured up to ((sin θ)/λ) ) 0.65 Å-1 on a Nonius KappaCCD instrument with rotating anode (graphite monochromator, Mo KR, λ ) 0.710 73 Å) at a temperature of 150 K. An analytical absorption correction was applied (µ ) 3.94 mm-1, 0.27-0.72 correction range). There were 7212 unique reflections (Rint ) 0.049). The structure was solved with automated Patterson methods with the program DIRDIF9926 and refined with the program SHELXL9727 against F2 of all reflections. Non-hydrogen atoms were refined freely with

Organometallics, Vol. 24, No. 21, 2005 5175 anisotropic displacement parameters. All hydrogen atoms were located in the difference Fourier map and were refined as rigid groups. The pentane solvent molecule is located on a 2-fold axis in the unit cell. There were 348 refined parameters, with 20 restraints. R (I > 2σ(I)): R1 ) 0.0237, wR2 ) 0.0545. R (all reflections): R1 ) 0.0321, wR2 ) 0.0586. GOF ) 1.065. The residual electron density was between -1.52 and 0.94 e/Å3. The drawings, geometry calculations, and checks for higher symmetry were performed with the program PLATON.28

Acknowledgment. This work was supported in part by the Council for Chemical Sciences of The Netherlands Organization for Scientific Research (NWO/CW). We thank Dr. M. Smoluch for the exact mass determination. Supporting Information Available: Tables giving Cartesian coordinates (Å) and energies (au) of all stationary points and a CIF file giving crystallographic data for compound 10. This material is available free of charge via the Internet at http://pubs.acs.org. OM050530+

(26) Beurskens, P. T.; Admiraal, G.; Beurskens, G.; Bosman, W. P.; Garcia-Granda, S.; Gould, R. O.; Smits, J. M. M.; Smykalla, C. The DIRDIF99 Program System; Technical Report of the Crystallography Laboratory, University of Nijmegen, Nijmegen, The Netherlands, 1999.

(27) Sheldrick, G. M. SHELXL-97: Program for Crystal Structure Refinement; Universita¨t Go¨ttingen, Go¨ttingen, Germany, 1997. (28) Spek, A. L. J. Appl. Crystallogr. 2003, 36, 7-13.